首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Fischer JC  Boughriet A  Wartel M 《Talanta》1981,28(4):233-236
The electrochemical reduction of NO(2)ClO(4) solutions at a platinum electrode in sulpholane has been studied. The results suggest the following scheme for the reduction of NO(+)(2) : 2NO(+)(2) + 2e(-) right harpoon over left harpoon 2NO(2) right harpoon over left harpoon N(2)O(4). A fast dimerization of nitrogen dioxide, which is favoured by the high donor number of sulpholane, follows the charge-transfer step. The NO(+)(2) ion is much more oxidative than the NO(+) ion in sulpholane, whereas the oxidizing power of the two species is nearly the same in nitromethane or acetonitrile solution. This result can be used to explain the role of water present at trace level.  相似文献   

2.
Issa IM  Idriss KA  Ghoneim MM 《Talanta》1976,23(3):249-251
The factors affecting the success of both visual and potentiometric end-point detection in titration of bisulphite with permanganate in the presence of fluoride are examined. The optimum conditions are 0.02M H(2)SO(4) and 0.24-0.38M NaF. The oxidation product comprises dithionate and sulphate according to the overall reaction MnO(4)(-) + H(+) + 2HF(2)(-) + 3HSO(3)(-) right harpoon over left harpoon MnF(4)(-) + S(2)O(6)(2-) + SO(4)(2-) + 3H(2)O. The reverse titration is also satisfactory, but proceeds quantitatively according to MnO(4)(2-) + 2HF(2)(-) + 2HSO(3)(-) right harpoon over left harpoon MnF(4)(-) + 2SO(4)(2-) + 2H(2)O.  相似文献   

3.
[MnIV3(micro-O)4(phen)4(H2O)2]4+ (, phen=1,10-phenanthroline) equilibrates with its conjugate base [Mn3(micro-O)4(phen)4(H2O)(OH)]3+ in aqueous solution. Among the several synthetic multinuclear oxo- and/or carboxylato bridged manganese complexes known to date containing metal-bound water, to the best of our knowledge, only deprotonates (right harpoon over left harpoon+H+, pKa=4.00 (+/-0.15) at 25.0 degrees C, I=1.0 M, maintained with NaNO3) at physiological pH. An aqueous solution of quantitatively oxidises NIII (HNO2 and NO2-) to NO3- within pH 2.3-4.1, the end manganese state being MnII. Both and are reactive oxidants in the title redox. In contrast to a common observation that anions react quicker than their conjugate acids in reducing metal centred oxidants, HNO2 reacts faster than NO2- in reducing or . The observed rates of nitrite oxidation do not depend on the variation of 1,10-phenanthroline content of the solution indicating that the MnIV-bound phen ligands do not dissociate in solution under experimental conditions. Also, there was no kinetic evidence for any kind of pre-equilibrium replacement of MnIV-bound water by nitrite prior to electron transfer which indicates the substitution-inert nature of the MnIV-bound waters and the 1,10-phenanthroline ligands. The MnIV3 to MnII transition in the present observation proceeds through the intermediate generation of the spectrally characterised mixed-valent MnIIIMnIV dimer that quickly produces MnII. The reaction rates are substantially lowered when solvent H2O is replaced by D2O and a rate determining 1e, 1H+ electroprotic mechanism is proposed.  相似文献   

4.
The kinetics of decomposition of peroxynitrous acid (ONOOH) was investigated in the presence of 0.1-0.75 M HClO(4) and at a constant ionic strength. The decay rate of ONOOH decreased in the presence of H(2)O(2), approaching a limiting value well below 75 mM H(2)O(2). It also decreased in the presence of relatively low [HNO(2)] but did not approach a lower limiting value, since ONOOH reacts directly with HNO(2). The latter reaction corresponds to a HNO(2)- and H(+)-catalyzed isomerization of ONOOH to nitrate, and its third-order rate constant was determined to be 520 +/- 30 M(-)(2) s(-)(1). The mechanism of formation of O(2)NOOH from ONOOH in the presence of H(2)O(2) was also scrutinized. The results demonstrated that in the presence of 0.1-0.75 M HClO(4) and 75 mM H(2)O(2) the formation of O(2)NOOH is insignificant. The most important finding in this work is the reversibility of the reaction ONOOH + H(2)O right harpoon over left harpoon HNO(2) + H(2)O(2), and its equilibrium constant was determined to be (7.5 +/- 0.4) x 10(-)(4) M. Using this value, the Gibbs' energy of formation of ONOOH was calculated to be 7.1 +/- 0.2 kcal/mol. This figure is in good agreement with the value determined previously from kinetic data using parameters for radicals formed during homolysis of peroxynitrite.  相似文献   

5.
The reaction equilibria [H(4)L](2+) + Zn(OAc)(2) right harpoon over left harpoon [Zn(H(2)L)](2+) + 2HOAc (K(1)) and [Zn(H(2)L)](2+) + Zn(OAc)(2) right harpoon over left harpoon [Zn(2)L](2+) + 2HOAc (K(2)), involving zinc acetate and the perchlorate salts of the tetraiminodiphenol macrocycles [H(4)L(1)(-)(3)](ClO(4))(2), the lateral (CH(2))(n)() chains of which vary between n = 2 and n = 4, have been studied by spectrophotometric and spectrofluorimetric titrations in acetonitrile. The photoluminescence behavior of the complexes [Zn(2)L(1)](ClO(4))(2), [Zn(2)L(2)(H(2)O)(2)](ClO(4))(2), [Zn(2)L(2)(mu-O(2)CR)](ClO(4)) (R = CH(3), C(6)H(5), p-CH(3)C(6)H(4), p-OCH(3)C(6)H(4), p-ClC(6)H(4), p-NO(2)C(6)H(4)), and [Zn(2)L(3)(mu-OAc)](ClO(4)) have been investigated. The X-ray crystal structures of the complexes [Zn(2)L(2)(H(2)O)(2)](ClO(4))(2), [Zn(2)L(3)(mu-OAc)](ClO(4)), and [Zn(2)L(2)(mu-OBz)(OBz)(H(3)O)](ClO(4)) have been determined. The complex [Zn(2)L(2)(mu-OBz)(OBz)(H(3)O)](ClO(4)) in which the coordinated water molecule is present as the hydronium ion (H(3)O(+)) on deprotonation gives rise to the neutral dibenzoate-bridged compound [Zn(2)L(2)(mu-OBz)(2)].H(2)O. The equilibrium constants (K) for the reaction [Zn(2)L(2)(H(2)O)(2)](2+) + A(-) right harpoon over left harpoon [Zn(2)L(2)A](+) + 2H(2)O (K), where A(-) = acetate, benzoate, or the carboxylate moiety of the amino acids glycine, l-alanine, l-histidine, l-valine, and l-proline, have been determined spectrofluorimetrically in aqueous solution (pH 6-7) at room temperature. The binding constants (K) evaluated for these systems vary in the range (1-8) x 10(5).  相似文献   

6.
The pulse radiolysis of aqueous NO has been reinvestigated, the variances with the prior studies are discussed, and a mechanistic revision is suggested. Both the hydrated electron and the hydrogen atom reduce NO to yield the ground-state triplet (3)NO(-) and singlet (1)HNO, respectively, which further react with NO to produce the N(2)O(2)(-) radical, albeit with the very different specific rates, k((3)NO(-) + NO) = (3.0 +/- 0.8) x 10(9) and k((1)HNO + NO) = (5.8 +/- 0.2) x 10(6) M(-)(1) s(-)(1). These reactions occur much more rapidly than the spin-forbidden acid-base equilibration of (3)NO(-) and (1)HNO under all experimentally accessible conditions. As a result, (3)NO(-) and (1)HNO give rise to two reaction pathways that are well separated in time but lead to the same intermediates and products. The N(2)O(2)(-) radical extremely rapidly acquires another NO, k(N(2)O(2)(-) + NO) = (5.4 +/- 1.4) x 10(9) M(-)(1) s(-)(1), producing the closed-shell N(3)O(3)(-) anion, which unimolecularly decays to the final N(2)O + NO(2)(-) products with a rate constant of approximately 300 s(-)(1). Contrary to the previous belief, N(2)O(2)(-) is stable with respect to NO elimination, and so is N(3)O(3)(-). The optical spectra of all intermediates have also been reevaluated. The only intermediate whose spectrum can be cleanly observed in the pulse radiolysis experiments is the N(3)O(3)(-) anion (lambda(max) = 380 nm, epsilon(max) = 3.76 x 10(3) M(-)(1) cm(-)(1)). The spectra previously assigned to the NO(-) anion and to the N(2)O(2)(-) radical are due, in fact, to a mixture of species (mainly N(2)O(2)(-) and N(3)O(3)(-)) and to the N(3)O(3)(-) anion, respectively. Spectral and kinetic evidence suggests that the same reactions occur when (3)NO(-) and (1)HNO are generated by photolysis of the monoprotonated anion of Angeli's salt, HN(2)O(3)(-), in NO-containing solutions.  相似文献   

7.
Isomerization of N(2)O(4) and dimerization of NO(2) in thin water films on surfaces are believed to be key steps in the hydrolysis of NO(2), which generates HONO, a significant precursor to the OH free radical in lower atmosphere and high-energy materials. Born-Oppenheimer molecular dynamics simulations using the density functional theory are carried out for NO(2)(H(2)O)(m), m ≤ 4, and N(2)O(4)(H(2)O)(n) clusters, n ≤ 7, used to mimic the surface reaction, to investigate the mechanism around room temperature. The results are (i) the NO(2) dimerization and N(2)O(4) isomerization reactions occur via two possible pathways, the non-water-assisted and water-assisted mechanisms; (ii) the NO(2) dimerization in the presence of water yields either ONONO(2)(H(2)O)(m) or NO(3)(-)NO(+)(H(2)O)(m) clusters, but it is also possible to form the HNO(3)(NO(2)(-))(H(3)O(+))(H(2)O)(m-2) transition state to form HONO and HNO(3), directly; (iii) the N(2)O(4) isomerization yields the NO(3)(-)NO(+)(H(2)O)(n) cluster, but it does not hydrolyze faster than the NO(2)(+)NO(2)(-)(H(2)O)(n) hydrolysis to directly form the HONO and HNO(3). New insights for hydrolysis of oxides of nitrogen in and on thin water films on surfaces in the atmosphere are discussed.  相似文献   

8.
Calculations were performed to determine the structures, energetics, and spectroscopy of the atmospherically relevant complexes (HNO(3)).(NO(2)), (HNO(3)).(N(2)O(4)), (NO(3)(-)).(NO(2)), and (NO(3)(-)).(N(2)O(4)). The binding energies indicate that three of the four complexes are quite stable, with the most stable (NO(3)(-)).(N(2)O(4)) possessing binding energy of almost -14 kcal mol(-1). Vibrational frequencies were calculated for use in detecting the complexes by infrared and Raman spectroscopy. An ATR-FTIR experiment showed features at 1632 and 1602 cm(-1) that are attributed to NO(2) complexed to NO(3)(-) and HNO(3), respectively. The electronic states of (HNO(3)).(N(2)O(4)) and (NO(3)(-)).(N(2)O(4)) were investigated using an excited state method and it was determined that both complexes possess one low-lying excited state that is accessible through absorption of visible radiation. Evidence for the existence of (NO(3)(-)).(N(2)O(4)) was obtained from UV/vis absorption spectra of N(2)O(4) in concentrated HNO(3), which show a band at 320 nm that is blue shifted by 20 nm relative to what is observed for N(2)O(4) dissolved in organic solvents. Finally, hydrogen transfer reactions within the (HNO(3)).(NO(2)) and (HNO(3)).(N(2)O(4)) complexes leading to the formation of HONO, were investigated. In both systems the calculated potential profiles rule out a thermal mechanism, but indicate the reaction could take place following the absorption of visible radiation. We propose that these complexes are potentially important in the thermal and photochemical production of HONO observed in previous laboratory and field studies.  相似文献   

9.
Daneţ AF  David V 《Talanta》1992,39(10):1299-1306
The partition constants of Cadion, i.e., 1-(p-nitro-phenyl)-3-(p'-azobenzene)-triazene, of its complex with the methylmercuric ion, and of methylmercury chloride were determined in the system toluene/aqueous phase containing 40 vol.% methyl alcohol; they have the values of 4.3 x 10(3), 3.0 x 10(3), and 2.6 respectively. The reagent has an absorption maximum at 406 nm, whereas the methylmercury complex at 460 nm. The K(HR) value corresponding to the H(+) + R(-) right harpoon over left harpoon HR equilibrium is 10(10.85), HR being the reagent molecule and H belongs to the NH of the triazenic group (NNNH). The K(ext) value corresponding to the equilibrium H(3)CHg(+) + (HR)(o) right harpoon over left harpoon (H(3)CHgR)(o) + H(+) is 1.0, where the "o" indicates the species present in the organic phase. The reagent/H(3)CHg(+) combination ratio is 1/1. The formation constant of the methylmercury complex, K(H(3)CHgR), which corresponds to the equilibrium H(3)CHg(+) + R(-) right harpoon over left harpoon H(3)CHgR, has a value of 10(10.8) as estimated by means of two different methods. The IR spectra allowed some conclusions to be drawn concerning the formation of the complex. The complex is stable up to 180 degrees , and the reagent up to 140 degrees . The molar absorptivity is of 3.46 x 10(4) 1.mole(-1).cm(-1) and the H(3)CHg(+) can be determined in the range 0.025-4 ppm. The determination is highly selective.  相似文献   

10.
Kinetics of the overall reaction [Cr(3)O(O(2)CCH(3))(6)(H(2)O)(3)](+) + 3 urea right harpoon over left harpoon [Cr(3)O(O(2)CCH(3))(6)(urea)(3)](+) + 3H(2)O have been studied spectrophotometrically. Monophasic kinetics were observed in both directions. The reverse steps, of urea dissociation, were monitored using an analytical technique which permits direct determination of the concentration of liberated urea and does not require knowledge of extinction coefficients of intermediate species. Results imply that consecutive steps occur with rate constants in close to the statistical ratios of k(1):k(2):k(3) = 3:2:1 and k(-)(1):k(-)(2):k(-)(3) = 1:2:3. Rates indicate strong labilization of urea, compared to the case of mononuclear complex [Cr(urea)(6)](3+).  相似文献   

11.
Xing W  Ingman F 《Talanta》1982,29(8):707-711
The complexation reaction between Alizarin complexan ([3-N,N-di(carboxymethyl)aminomethyl]-1,2-dihydroxyanthraquinone; H(4)L) and zinc(II), nickel(II), lead(II), cobalt(II) and copper(II) has been studied by a spectrophotometric method. All these metal ions form 1:1 complexes with HL; 2:1 metal:ligand complex were found only for Pb(II) and Cu(II). The stability constants are (ionic strength I = 0.1, 20 degrees C): Zn(2+) + HL(3-) right harpoon over left harpoon ZnHL(-) log K +/- 3sigma(log K) = 12.19 +/- 0.09 (I = 0.5) Ni(2+) + HL(3-) right harpoon over left harpoon NiHL(-) log K +/- 3sigma(log K) = 12.23 +/- 0.21 Pb(2+) + HL(3-) right harpoon over left harpoon PbHL(-) log K +/- 3sigma(log K) = 11.69 +/- 0.06 PbHL(-) + Pb(2+) right harpoon over left harpoon Pb(2)L + H(+) log K approximately -0.8 Co(2+) + HL(3-) right harpoon over left harpoon CoHL(-) log K 3sigma(log K) = 12.25 + 0.13 Cu(2+) + HL(3-) right harpoon over left harpoon CuHL(-) log K 3sigma(log K) = 14.75 +/- 0.07 Cu(2+) + CuHL(-) right harpoon over left harpoon Cu(2)L + H(+) log K approximately 3.5 The solubility and stability of both the reagent and the complexes and the closenes of the values of the stability constants make this reagent suitable for the photometric detection of several metal ions in the eluate from an ion-exchange column.  相似文献   

12.
Pesavento M 《Talanta》1989,36(10):1059-1064
The potential of the As(V)/As(III) half-cell was measured at 25 degrees with a glass electrode as reference electrode in order to eliminate the liquid-junction potential. Rapid and reproducible values could be obtained only in the presence of iodide, which increases the rate of electron-exchange between the two oxidation states of arsenic, but only at hydrogen-ion concentrations higher than about 0.5M. Extrapolation to zero ionic strength was therefore required to obtain the standard potential. A value of 573 +/- 2 mV was calculated for the half-reaction AsO(OH)(3) + 2e(-) + 2H(+) right harpoon over left harpoon As(OH)(3) + H(2)O.  相似文献   

13.
The title reaction takes place according to the stoichiometry 2L(2)RhOO(2+) + 3HNO(2) + H(2)O --> 2L(2)Rh(OH(2))(3+) + 3NO(3)(-) + H(+) (L(2) = meso-Me(6)-[14]ane-N(4)). The kinetics are second order in HNO(2) and independent of the concentration of L(2)RhOO(2+), rate = (k(1) + k(2)[H(+)])[HNO(2)](2), where k(1) = 10.9 M(-1) s(-1) and k(2) = 175 M(-2) s(-1) at 25 degrees C and 0.10 M ionic strength. The reaction produces two observable intermediates, the nitrato (L(2)RhONO(2)(2+)) and hydroperoxo (L(2)RhOOH(2+)) complexes. The product analysis and kinetics are indicative of the initial rate-controlling formation of NO and NO(2), both of which react rapidly with L(2)RhOO(2+) in subsequent steps. The reaction with NO produces mainly L(2)RhONO(2)(2+), which hydrolyzes to L(2)Rh(OH(2))(3+) and NO(3)(-). Another minor pathway generates the hydroperoxo complex, which was detected by its known reaction with Fe(aq)(2+). The reaction of NO(2) with L(2)RhOO(2+) requires an additional equivalent of HNO(2) and produces L(2)Rh(OH(2))(3+) and NO(3)(-) via a proposed peroxynitrato complex L(2)RhOONO(2)(2+). This work provides strong evidence for the long-debated reaction between HNO(2) and H(2)NO(2)(+) to generate N(2)O(3).  相似文献   

14.
The kinetics and mechanism of the chlorine(III)-HOBr reaction were studied by the stopped-flow method under acidic conditions, pH 1.0-3.0, in 1.0 M NaClO(4) and at 25.0 degrees C. The overall redox process occurs in two consecutive steps via the formation of the BrClO(2) intermediate. The electron transfer reactions are coupled with bromine hydrolysis, the formation of the tribromide ion, and the protolytic equilibrium of chlorine(III). On the basis of simultaneous evaluation of the kinetic traces, the following rate constants were obtained for the redox steps: HClO(2) + HOBr right harpoon over left harpoon BrClO(2) + H(2)O, k(3) = (3.34 +/- 0.02) x 10(4) M(-1) s(-1), k(-3) = (3.5 +/- 1.3) x 10(3) s(-1); BrClO(2) + ClO(2)(-)<==>2ClO(2) + Br(-), k(4) = (2.9 +/- 1.0) x 10(7) M(-1) s(-1). The second step was practically irreversible under the conditions applied, and the value of k(-4) could not be determined. The equilibrium constant for the formation of BrClO(2), K(3) = 9.5 M(-1), was calculated from the kinetic results, and it was confirmed that this species is a very powerful oxidant. The redox potential was also estimated for the BrClO(2) + e(-) = Br(-) + ClO(2) reaction: epsilon(0) approximately 1.70 V.  相似文献   

15.
Silyl anion SiButPh2- (2) was found to substitute an amide ligand in Zr(NMe2)4 (3) to give the disilyl complex Zr(NMe2)3(SiButPh2)2- (1a) and Zr(NMe2)5- (1b) in THF. The reaction is reversible, and nucleophilic amide NMe2- attacks the Zr-SiButPh2 bonds in 1a or Zr(NMe2)3(SiButPh2) in the reverse reaction, leading to an unusual ligand exchange equilibrium 2 3 + 2 2 right harpoon over left harpoon 1a + 1b (eq 1). The silyl anion 2 selectively attacks the -N(SiMe3)2 ligand in Zr(NMe2)3[N(SiMe3)2] (6) to give 1a and N(SiMe3)2- (7). Reversible reaction occurs as well, where 7 selectively substitutes the silyl ligand in Zr(NMe2)3(SiButPh2)2- (1a) or Zr(NMe2)3(SiButPh2), giving the equilibrium 6 + 2 2 right harpoon over left harpoon 1a + 7 (eq 3). The thermodynamics of these equilibria has been studied: For eq 1, DeltaH degrees = -8.3(0.2) kcal/mol, DeltaS degrees = -23.3(0.9) eu, and DeltaG degrees 298K = -1.4(0.5) kcal/mol at 298 K; for eq 3, DeltaH degrees = -1.61(0.12) kcal/mol, DeltaS degrees = -2.6(0.5) eu, and DeltaG degrees 298K = -0.8(0.3) kcal/mol. In both equilibria, the enthalpy changes for the forward reactions outweigh the entropy changes, and therefore the substitutions of amide ligands in Zr(NMe2)4 (3) and Zr(NMe2)3[N(SiMe3)2] (6) to afford the disilyl complex 1a are thermodynamically favored. The following equilibria were also observed and studied: Zr(NMe2)3[N(SiMe3)2] (6) + Si(SiMe3)3- (9) right harpoon over left harpoon Zr(NMe2)3[Si(SiMe3)3] (10) + N(SiMe3)2- (7) and Zr(NMe2)4 (3) + 9 right harpoon over left harpoon 10 + Zr(NMe2)5- (1b).  相似文献   

16.
Chen MH  Lee S  Liu S  Yeh A 《Inorganic chemistry》1996,35(9):2627-2629
Reactions of Fe(CN)(5)L(3-) (L = 4-aminopyridine (4-ampy), pyridine (py), 4,4'-bipyridine (4,4'-bpy), and pyrazine (pz)) with peroxydisulfate, Fe(CN)(5)L(3-) + S(2)O(8)(2-) right harpoon over left harpoon Fe(CN)(5)L(2-) + SO(4)(-) + SO(4)(2-), have been found to follow an outer-sphere electron transfer mechanism. The specific rate constants of oxidation are 1.45 +/- 0.01, (9.00 +/- 0.02) x 10(-2), (5.60 +/- 0.01) x 10(-2), and (2.89 +/- 0.01) x 10(-2) M(-1) s(-1), for L = 4-ampy, py, 4,4'-bpy, and pz, respectively, at &mgr; = 0.50 M LiClO(4), T = 25 degrees C, pH = 4.4-8.8. The rate constants of oxidation for the corresponding Ru(NH(3))(5)L(2+) complexes were also measured and were found to be faster than those of Fe(CN)(5)L(3-) complexes by a factor of approximately 10(2) even after the corrections for the differences in reduction potentials and in the charges of the complexes. The difference in reactivity may arise from the hydrogen bonding between peroxydisulfate and the ammonia ligands of Ru(NH(3))(5)L(2+) and nonadiabaticity observed in the Fe(CN)(5)L(3-) complexes.  相似文献   

17.
Raman spectra of HNO(3).NO(2) have been detected on liquid and solid surfaces in the presence of concentrated HNO(3) and NO(2) gas. The Raman spectrum of HNO(3) solutions containing N(2)O(4) has been partly reinterpreted in terms of contributions from HNO(3).N(2)O(4) and N(2)O(4).NO(3)(-) complexes.  相似文献   

18.
Equilibria in the U(VI)H(2)OCO(2)(g) system in 0.5M sodium perchlorate medium at 25 degrees have been studied. By using thermal tensing spectrophotometry (TLS) and a very low total concentration of U(V1) (4 x 10(-6)M) information could be obtained on equilibria involving UO(2)(CO(3))(2-)(2) without complications due to formation of the trimer (UO(2))(3)(CO(3))(6-)(6). The experimental data allowed a precise determination of the equilibrium constant log K(3) = 6.35 +/- 0.05 for the reaction UO(2)(CO(3))(2-)(2) + CO(2-)(3) right harpoon over left harpoonright harpoon over left harpoon UO(2)(CO(3))(4-)(3). The interpretation of TLS data is briefly discussed, as well as the potential use of this technique for studies of the speciation of trace elements in natural water systems.  相似文献   

19.
The first mechanistic study of a spin-forbidden proton-transfer reaction in aqueous solution is reported. Laser flash photolysis of alkaline trioxodinitrate (N(2)O(3)(2)(-), Angeli's anion) is used to generate a nitroxyl anion in its excited singlet state ((1)NO(-)). Through rapid partitioning between protonation by water and electronic relaxation, (1)NO(-) produces (1)HNO (ground state, yield 96%) and (3)NO(-) (ground state, yield 4%), which comprise a unique conjugate acid-base couple with different ground-state multiplicities. Using the large difference between reactivities of (1)HNO and (3)NO(-) in the peroxynitrite-forming reaction with (3)O(2), the kinetics of spin-forbidden deprotonation reaction (1)HNO + OH(-) --> (3)NO(-) + H(2)O is investigated in H(2)O and D(2)O. Consistent with proton transfer, this reaction exhibits primary kinetic hydrogen isotope effect k(H)/k(D) = 3.1 at 298 K, which is found to be temperature-dependent. Arrhenius pre-exponential factors and activation energies of the second-order rate constant are found to be: log(A, M(-)(1) s(-)(1)) = 10.0 +/- 0.2 and E(a) = 30.0 +/- 1.1 kJ/mol for proton transfer and log(A, M(-)(1) s(-)(1)) = 10.4 +/- 0.1 and E(a) = 35.1 +/- 0.7 kJ/mol for deuteron transfer. Collectively, these data are interpreted to show that the nuclear reorganization requirements arising from the spin prohibition necessitate significant activation before spin change can take place, but the spin change itself must occur extremely rapidly. It is concluded that a synergy between the spin prohibition and the reaction energetics creates an intersystem barrier and is responsible for slowness of the spin-forbidden deprotonation of (1)HNO by OH(-); the spin prohibition alone plays a minor role.  相似文献   

20.
Issa IM  Ghoneim MM 《Talanta》1973,20(6):517-524
The effects of acidity, fluoride concentration, temperature and concentration of manganese in the reaction between KMnO(4) and Mn(II) were studied potentiometrically. The rate of reaction is increased by increasing the fluoride concentration and/or decreasing the acidity of the solution. The formal redox potentials of the MnO(-)(4)/Mn(III) and the Mn(III)/Mn(II) systems were determined at different pH values. The E degrees values obtained by extrapolation to pH = 0 were 1.58 and 1.52 V respectively. The amount of Mn(II) determined was varied from 5 to 56 mg. The net reaction can be represented as MnO(-)(4) + 10HF(-)(2) + 4Mn(2+) right harpoon over left harpoon 5MnF(-)(4) + 2H(+) + 4H(2)O.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号